Next Article in Journal
Ultrasound/Chlorine: A Novel Synergistic Sono-Hybrid Process for Allura Red AC Degradation
Next Article in Special Issue
Development of Silicon Carbide-Supported Palladium Catalysts and Their Application as Semihydrogenation Catalysts for Alkynes under Batch- and Continuous-Flow Conditions
Previous Article in Journal
Photocatalytic Degradation of Methylene Blue under Visible Light Using TiO2 Thin Films Impregnated with Porphyrin and Anderson-Type Polyoxometalates (Cu and Zn)
Previous Article in Special Issue
Mononuclear Oxidovanadium(IV) Complexes with BIAN Ligands: Synthesis and Catalytic Activity in the Oxidation of Hydrocarbons and Alcohols with Peroxides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copper Supported on MgAlOx and ZnAlOx Porous Mixed-Oxides for Conversion of Bioethanol via Guerbet Coupling Reaction

1
Hangzhou Institute of Advanced Studies, Zhejiang Normal University, Hangzhou 311231, China
2
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
3
The State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(10), 1170; https://doi.org/10.3390/catal12101170
Submission received: 30 August 2022 / Revised: 26 September 2022 / Accepted: 27 September 2022 / Published: 4 October 2022
(This article belongs to the Special Issue Feature Papers in Catalysis in Organic and Polymer Chemistry)

Abstract

:
The direct conversion of biomass-derived ethanol to high-valued-added chemicals has attracted widespread attention recently due to the great economic and environmental advantages. In the present study, the conversion of bioethanol through the Guerbet coupling process was studied in a fixed-bed reactor for MgAlOx and ZnAlOx mixed-oxides supported Cu catalysts. From the results, Cu adding into the system greatly enhance the dehydrogenation of ethanol and increase the H-transfer in the course of Guerbet coupling process. Simultaneously, the porous mixed-oxides provide the acid-base property of the catalysts for intermediate transformation. Notably, for Cu/MgAlOx, the main product of ethanol conversion is butanol, but for Cu/ZnAlOx, the primary product is ethyl acetate. Characterizations such as X-ray diffraction (XRD), high-resolution transmission electron microscopy (HRTEM), X-ray photoelectron spectroscopy (XPS), in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and CO2 temperature programmed desorption (TPD) were carried out to evaluate the structure and property of the catalysts. In combination with the catalytic performances with the characterization results, the synergistic catalytic effect between metal sites and acid-base sites were elaborated.

1. Introduction

Over the past decades, due to the increased depletion of fossil feedstock, the demand for alternate and renewable energy is continuously growing [1,2]. Ethanol is a versatile and sustainable raw material which can be produced from the fermentation of renewable biomass such as sugars and corns [3,4]. Currently, bioethanol is widely used as a fuel additive, blending with gasoline for partly replacing traditional fossil fuels [5,6]. However, when compared with gasoline, ethanol has some major drawbacks, such as low energy density (19.6 MJ·L−1), high water solubility, and corrosion compared to the current technology of engines and fuel infrastructure [7,8,9,10]. Compared to ethanol, butanol has larger energy density (29.2 MJ·L−1) and lower miscibility in water. Thus, it is considered as a potential gasoline fuel additive and “advanced biofuel” with good environmental benefits [8,9]. In addition to use as a biofuel, butanol can be also utilized as an important raw bulk material in the manufacture of paints, solvents, and plasticizers [11,12].
Traditionally, n-butanol is produced through fossil-based oxo process or the fermentation of sugar-containing crops (ABE process) [13,14]. Alternatively, with the increased availability of ethanol from biomass, the direct conversion of ethanol into butanol has been proposed to be an economical and sustainable route for butanol production [15,16,17,18]. The Guerbet coupling reaction, corresponding to the catalytic conversion of light alcohols into higher ones, have recently attracted widespread attention [19,20]. Several heterogeneous catalysts such as hydroxyapatites (HAP) [21,22], metal oxides [8,12,23], activated carbon [17,24,25], and molecular sieves [26] were utilized to convert ethanol into higher alcohols through the Guerbet coupling process. However, under the present catalytic systems, the reaction condition needs to be performed at high temperatures (generally above 623 K) [12,21,26]. Thus, many researchers have tried their best to decrease the operating temperature and increase the activity of catalysts by adding metal species [14,27,28,29,30].
Hydrotalcite (HT) derived mixed oxides supported copper catalysts have been found to be a good candidate for the Guerbet coupling process due to the synergistic effect between metal species and mixed oxides by coupling dehydrogenation/hydrogenation process and condensation reactions for ethanol transformation [23,27,28]. For example, Cu/MgxAlOy catalysts with low Cu loadings (0.1–0.6 wt.%) exhibited high selectivity (49–63%) to linear chain C4+ alcohols [28]. Cu has been proven to have a promotional effect on the reaction rates of H-transfers, accelerating acid-base-catalyzed deprotonation and hydrogenation/dehydrogenation steps. Cu modified NiMgAlO catalysts were reported to improve ethanol conversion and butanol selectivity at moderate reaction conditions (523 K) [27]. The presence of Cu species was supposed to create Lewis acid-base pairs and CuNi alloy sites, thus increasing H-transfer and condensation reactions, resulting in elevated ethanol conversion (30%) and butanol selectivity (64.2%). Recently, our group found that Cu supported on NiAlOx exhibited good performance and stability in the Guerbet coupling process [31]. An optimal compromise between the ethanol conversion (~35%) and butanol selectivity (~45%) has to be sustained for the 0.75%Cu/NiAlOx catalyst for 1000 h at 523 K [31]. However, for the above catalysts, the effect of support was seldomly elaborated, which in most cases, influencing the property and performances of the catalysts [32,33,34,35]. Additionally, the product distribution of the Guerbet coupling process was highly related with the support effect when synthesized with other valuable chemicals such as acetaldehyde, ethyl acetate, ether, butadiene, etc. [32,35].
This research aims to investigate the MgAlOx and ZnAlOx porous mixed-oxides supported Cu catalysts for ethanol conversion to butanol via Guerbet coupling reaction. With similar synthesis process and Cu contents, the catalytic performance and stability of both catalysts were studied systematically. N2-physical adsorption-desorption, X-ray diffraction (XRD), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy (XPS) were carried out to evaluate the geometric structure and electronic property of the catalysts. Additionally, CO2 temperature programmed desorption (TPD), pyridine adsorbed Fourier-transformed infrared absorption spectra (FT-IR), and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) were employed to test the acid-base property and adsorptive property of the catalysts, from which the relationship between the property and catalytic performances were accordingly revealed.

2. Results and Discussion

2.1. Catalytic Performances

The catalytic performance of MgAlOx and ZnAlOx porous mixed-oxides supported Cu catalysts for ethanol conversion were carried out in a continuous fixed-bed reactor at 553 K and 2 MPa of N2. For comparison, pure mixed-oxides without addition of Cu were carried out for this reaction under the identical reaction conditions. From Table 1, MgAlOx and ZnAlOx (entries 1 and 2) showed very low activity for ethanol, with values of 4.4% and 8.5%, respectively. In addition, the selectivity of butanol was very low for the above mixed-oxides. Nevertheless, when Cu was added into the mixed-oxides, ethanol conversion was greatly enhanced to 43.1% for Cu/MgAlOx (Table 1, entry 3) and 33.9% for Cu/ZnAlOx (Table 1, entry 4). Simultaneously, the selectivity of butanol for Cu/MgAlOx greatly increased to 33.2%, accomplished by major byproducts of ethyl acetate (Sel. 11.8%), butaldehyde (Sel. 8.6%), ethyl butyrate (Sel. 5.0%), and hexanol (Sel. 8.0%). Such results are comparable or better than the previous results, in which a good compromise between ethanol conversion and butanol selectivity was pursued [23,31,34]. In contrast, the main product on the Cu/ZnAlOx was ethyl acetate, with the selectivity of 42.2%, which is widely used in paints, coatings, inks, and adhesives. Other byproducts, such as butanol (Sel. 7.7%), ether (Sel. 1.8%), butaldehyde (Sel. 9.4%) and ethyl butyrate (Sel. 8.3%), can be also detected. Thus, the results suggested Cu species could facilitate the transformation of ethanol in the Guerbet coupling process, both in the Cu/MgAlOx and Cu/ZnAlOx catalytic systems. However, the support effect also makes a big difference on the distribution of products, implying a synergistic effect between Cu and mixed-oxides on the Guerbet coupling process.
To further evaluate the products distribution and stability of above catalysts, a time on stream (TOS) of above 100 h was performed for the Cu/MgAlOx and Cu/ZnAlOx catalysts for the Guerbet coupling process. The results are displayed in Figure 1. From Figure 1a, the major product of Cu/MgAlOx is butanol, with selectivity preserving in the range of 30~33%. Other byproducts including ethyl acetate, butane, butaldehyde, hexanol, ethyl butyrate, etc. can be also detected. However, the selectivities were all below 10%. Undetectable gas byproducts or liquid products might also be formed in the process, which account for 20% of the carbon mass loss. In addition, Cu/MgAlOx catalyst exhibited an obvious deactivation feature, with ethanol conversion decreased from 43% to 33% after a TOS of 108 h. Unlike the Cu/MgAlOx, the main product of the Cu/ZnAlOx is ethyl acetate, with the selectivity maintained at 40~42% (Figure 1b). Butane, butanol, ethyl butyrate, etc. appeared as the main byproducts. Except for a slight decrease in initial 24 h, the conversion of ethanol for Cu/ZnAlOx is preserved in the range of 29~33%, implying an excellent stability of Cu/ZnAlOx catalyst for ethanol conversion.

2.2. Structure of the Catalysts

N2-physical adsorption-desorption was evaluated to test the pore structure of catalysts such as the specific surface area (SBET), pore volume, and pore width (as displayed in Table 2). The isotherms and BJH pore size distribution curves are displayed in Figure 2. From the results, MgAlOx and ZnAlOx showed SBET of 119.7 and 155.9 m2/g, respectively. However, as Cu species was added, a decrease in SBET and pore volume were observed in Cu/MgAlOx and Cu/ZnAlOx catalysts, which could be attributed to the deposition of Cu species inside the pores of mixed oxides. Furthermore, the isothermal curves of above samples show obvious type IV hysteresis loops according to IUPAC classification, indicating the presence of porous structure in the above samples [36,37]. Additionally, the contents of Cu for the Cu/MgAlOx and Cu/ZnAlOx catalysts were measured by inductively coupled plasma atomic emission spectroscopy (ICP-AES), with the similar values of 1.11 wt.% and 1.35 wt.%, which also agreed well with the nominal values.
The powder XRD patterns of Cu/MgAlOx and Cu/ZnAlOx catalysts are displayed in Figure 3. From the results, both catalysts show characteristic peaks of corresponding oxides (MgO or ZnO), which derived from the collapse of layered hydrotalcite structure. The absence of Al2O3 phase in the two catalysts indicated an amorphous state of Al2O3 or homogeneous dispersion of Al2O3 inside the mixed oxides. In addition, in the Cu/MgAlOx and Cu/ZnAlOx catalysts, no diffraction peaks ascribed to Cu nanoparticles could be observed, indicating a uniformed dispersed Cu species or low concentration of Cu on the surface.
In order to get the structural information and lattice parameters of above catalysts, TEM and HRTEM were performed on the Cu/MgAlOx (Figure 4a,c) and Cu/ZnAlOx (Figure 4b,d) catalysts. From TEM images (Figure 4a,b), no obvious particles could be observed in the above samples, implying Cu species were highly dispersed on the surface of the catalysts. In addition, the interplanar spacing of ~2.14 Å and ~2.86 Å corresponding to MgO (200) and ZnO (100) lattice fringes could be discerned for Cu/MgAlOx and Cu/ZnAlOx (Figure 4d), implying the major phase of above catalysts is corresponding oxides, which was consistent with the XRD results.

2.3. Electronic Property of the Catalyst

The surface composition of Cu/MgAlOx and Cu/ZnAlOx catalysts were measured by X-ray photoelectron spectroscopy (XPS). The C 1s, Cu 2p, O 1s, and Al 2p XPS spectra are shown in Figure 5. The binding energies (B.E.) of the above elements were referenced; the B.E. of C 1s core level at 284.8 eV (Figure 5a) [38]. Figure 5b shows the Cu 2p XPS spectra of Cu/MgAlOx and Cu/ZnAlOx catalysts. A couple of peaks located at 933.0 eV and 952.8 eV were ascribed to Cu 2p3/2 and Cu 2p1/2, respectively [27,39]. The Cu 2p3/2 at 933.0 eV and 934.8 eV should be assigned to Cu+/Cu0 and Cu2+ species [31,39]. Noteworthy, the Cu species of the Cu/MgAlOx are almost Cu+/Cu0, whereas for the Cu/ZnAlOx, except for Cu+/Cu0, Cu2+ species also existed on the surface, as proved by the coexistence of satellite peaks at 940~945 eV [39,40]. Figure 5c shows the O 1s XPS spectra of the two catalysts. From the results, two distinct peaks at 531.6 eV and 530.2 eV were observed and should be ascribed to the existence of adsorbed oxygen (Oα) and lattice oxygen (Oβ) [27,41]. The relative amounts of Oα and Oβ changed with the composition of supports, suggesting that the chemical form of surface oxygen depends heavily on the chemical composition. Obviously, the Cu/MgAlOx has more Oα than that of the Cu/ZnAlOx, indicating that oxygen molecules prefer to accumulate on the surface of the Cu/MgAlOx in the form of a hydroxyl group, which might provide some basic sites to the catalysts [27]. In contrast, for the Cu/ZnAlOx catalyst, more Oβ were present on the surface, suggesting that oxygen appears in the form of metal oxides, i.e., ZnO or Al2O3. In addition, in the Al 2p XPS spectra (Figure 5d), the binding energy at 73.8–74.2 eV could be assigned to Al3+, which corresponded to the Zn-O-Al and Mg-O-Al, respectively [41].

2.4. The Acid-Basic Property of the Catalyst

Figure 6 shows the CO2 temperature-programmed desorption (TPD) profiles and pyridine adsorbed FT-IR spectra of the supported Cu catalysts. From the results, the acid-basic property of the Cu/MgAlOx and Cu/ZnAlOx catalysts were thereby compared. In Figure 6a, the surface basicity has been determined by CO2-TPD. In the range of 100~900 °C, the CO2 desorption peaks can be deconvoluted into four contributions, which could be assigned to the weak (<250 °C), moderate (250~470 °C), strong (470~650 °C), and super strong basic sites (>650 °C), respectively [42,43]. For Cu/MgAlOx, a large amount of moderate basic sites can be observed in CO2-TPD profiles, and the intensity is 10 times higher than that of Cu/ZnAlOx. Based on O 1s XPS spectra (Figure 5b), the moderate basic sites might derive from the surface hydroxyl group [33,43], whereas for the Cu/ZnAlOx, more lattice oxygen contributes to the strong and super strong basic sites [33,43]. In order to determine surface acidity, in situ FT-IR of pyridine adsorbed on Cu/MgAlOx and Cu/ZnAlOx were measured separately. The distribution of Brønsted and Lewis acid sites were determined in the range of 1300~1700 cm−1. As seen in Figure 6b, the absence of a band positioned at 1540 cm−1 in Cu/MgAlOx and Cu/ZnAlOx reflects the lack of Brønsted acid sites (B), whereas the band at 1450 cm−1 corresponding to pyridine chemisorbed on Lewis acid sites (L) could be clearly observed, indicating that only Lewis acidity is present [35,44]. Notably, the relative intensity of adsorption band at 1450 cm−1 is stronger on Cu/ZnAlOx than that of Cu/MgAlOx, suggesting that higher acidity is observed for the Cu/ZnAlOx catalyst.

2.5. In Situ Drift Spectra of Ethanol Adsorption and Transformation

In situ DRIFT spectra of ethanol adsorption and transformation for the Cu/MgAlOx and Cu/ZnAlOx catalysts were performed and are displayed in Figure 7. With the temperature increasing from 298 K to 573 K, ethanol gradually converted into different intermediates and products for the above catalysts, directely reflecting the reaction of the Guerbet coupling process. For Cu/MgAlOx (Figure 7a), the bands at 1252 cm1 are observed at 298 K, which is assigned to the δ(C-OH) of adsorbed 3-hydroxybutanal [45], whereas for Cu/ZnAlOx (Figure 7b), the vibration at 1393 cm1 could be discerned, indicating the presence of ethyl-acetate species [46]. In addition, the bands at 1076 and 1102 cm1 were presented in both Cu/MgAlOx and Cu/ZnAlOx catalysts, corresponding to C-O stretching vibrations in adsorbed ethoxide [27,41]. The Cu/ZnAlOx catalyst has higher intensity of adsorbed ethoxide at the elevated temperatures, implying that it is more difficult for ethanol transformation for Cu/ZnAlOx than for Cu/MgAlOx. The results were in accordance with the catalytic performance (Table 1 and Figure 1). Additionally, as the test temperature increased from 298 K to 523 K, two new bands positioned at 1645 and 1754 cm1 were produced, which is due to the stretching vibrations of C=C group and C=O group in adsorbed crotonaldehyde [27,47]. Furthermore, with the reaction temperature further increased to 573 K, the peaks at 1645 and 1754 cm1 become more prominent for Cu/MgAlOx, implying the intermediate of crotonaldehyde is easier to produce on Cu/MgAlOx on Cu/ZnAlOx, and probably leads to higher yields of a major product such as butanol (Table 1).

2.6. Insight into the Catalytic Performances

According to the catalytic performances on the MgAlOx and ZnAlOx supported Cu catalysts and the corresponding mixed oxides for ethanol transformation (Table 1 and Figure 1), the active sites were supposed to relate with the Cu species and the supports. Firstly, the structure of Cu species was evaluated by XRD (Figure 3) and TEM (Figure 4) techniques. From the results, Cu species were found highly dispersed on the MgAlOx and ZnAlOx supports before catalytic performance. After the catalytic test, Cu particles aggregated to larger ones, with the average particle sizes of 6.4 nm for Cu/MgAlOx and 4.8 nm for Cu/ZnAlOx, respectively (Figure S1). The increased metal particles might account for the decreased activity of both catalysts, especially for Cu/MgAlOx. EDX-mapping (Figure S2) and XRD results (Figure S3) indicated the Cu species were still highly dispersed on the catalysts, despite the increased particle sizes.
The valence states of Cu species in the Cu/MgAlOx and Cu/ZnAlOx catalysts before catalytic tests are mainly Cu0 and Cu+, as displayed by XPS (Figure 5). For Cu/ZnAlOx, except for Cu+/Cu0, Cu2+ species also existed on the surface. Figure S4 shows the surface composition of Cu/MgAlOx and Cu/ZnAlOx catalysts after stability tests. The Cu 2p XPS spectra (Figure S4c) indicated the Cu species are all Cu0/Cu+ on Cu/MgAlOx and Cu/ZnAlOx, as the satellite peaks attributed to Cu2+ disappeared after catalytic tests. Recently, many researches have reported the supported copper catalysts for the Guerbet coupling process [23,24,27,28,29]. Cu species in the form of Cu0/Cu+ were proved to have a promotional effect on the rates of reaction steps that involve H-transfers [27,29]. Currently, our group also found that the addition of Cu to the NiAlOx could greatly enhance the activity and selectivity of ethanol coupling to butanol [31]. A small amount of Cu0/Cu+ was sufficient for ethanol transformation and to improve the reaction rates of ethanol dehydrogenation [31]. In this work, Cu species with the valence state of Cu0/Cu+ were also considered to be the active sites for H-transfers, because the majority of Cu species for the Cu/MgAlOx and Cu/ZnAlOx catalysts before and after stability tests are Cu0 and Cu+.
Except for Cu species, the acid-base property resulting from different mixed oxides also plays an important role on the Guerbet coupling process. Previously, large amounts of works presented the conclusions that metal active sites benefit for H-transfers, and acidic-basic sites are in favor of the aldol condensation [27,28,29,30]. In this work, based on the results of CO2-TPD and pyridine adsorbed FT-IR spectra, the Cu/MgAlOx catalyst exhibits more basic sites, whereas the Cu/ZnAlOx catalyst shows more Lewis acid sites (Figure 6). Generally, basic metal oxides were reported active for catalyzing the aldol condensation reaction [44,45], while metal cations as Lewis acid normally in conjugation with basic sites form acid-base pairs for the adsorption and activation of acetaldehyde [48]. Hence, in the Guerbet coupling process, ethanol is dehydrogenated to acetaldehyde for the metal active site, followed by a base-catalyzed aldol coupling reaction to form 3-hydroxybutyraldehyde, then 3-hydroxybutyraldehyde dehydration to crotonaldehyde, and finally hydrogenation to butanol (Scheme 1a) [28,48]. For Cu/MgAlOx, sufficient basic sites made it a good candidate for production of butanol. Nevertheless, in synthesis of ethyl acetate, ethanol is firstly dehydrogenated to form acetaldehyde, then acetaldehyde is condensed with ethanol to form hemiacetal, and finally the hemiacetal dehydrogenated to form ethyl acetate (Scheme 1b) [14,49]. As Lewis acid sites can benefit for the adsorption and activation of acetaldehyde, more ethyl acetate would be generated on the acidic Cu/ZnAlOx catalyst [14,32,48]. In situ drifts of ethanol (Figure 7) reveal the dynamic reaction process in the Cu/MgAlOx and Cu/ZnAlOx catalysts, which is in agreement with the catalytic performance. The above results strongly suggest the synergic effect between Cu and mixed metal oxides plays an exclusive role in determining the products distribution in the Guerbet coupling reaction.

3. Materials and Methods

3.1. Chemicals

Copper nitrate trihydrate (Cu(NO3)2·3H2O, >99.5%), magnesium nitrate hexahydrate (Mg(NO3)2·6H2O, 99%), zinc nitrate hydrate ((Zn(NO3)2·6H2O), 99 wt%), and aluminum nitrate hexahydrate (Al(NO3)3·9H2O, 99%) were purchased from Sinopharm Chemical Reagent Co. LTD (Shanghai, China). Ethanol (≥99.8%), butanol (>99.5%), methanol (≥99.5%), sodium hydroxide (NaOH, ≥96.0%), sodium carbonate (Na2CO3, ≥99.8%), ethyl acetate (≥99.7%), and ortho-xylene (≥99.5%) were purchased from Aladdin Industrial Corporation (Shanghai, China). All chemicals were used as received and without any purification. All glassware was washed with Aqua Regia and rinsed with ethanol and ultrapure water. Ultrapure water (18.2 MΩ) was used throughout this work.

3.2. Preparation of the Supported Cu Catalysts

Hydrotalcite (HT) supports of MgAl-HT and ZnAl-HT were synthesized through coprecipitation method with Mg(Zn)/Al molar ratio of 3. In a typically synthesis, a certain amount of Mg(NO3)2·6H2O or Zn(NO3)2·6H2O (0.21 mol) and Al(NO3)3·9H2O (0.07 mol) were dissolved in 200 mL of deionized water to get solution A. Then, a certain amount of NaOH (0.438 mol) and Na2CO3 (0.113 mol) were dissolved in 200 mL of deionized water to get solution B. Under vigorous stirring at 75 °C, solution A was dropwise added into solution B with a constant flow of 3 mL/min, and then the mixture was aged at 75 °C for 24 h. The obtained suspension was filtered and washed for several times with deionized water, and dried overnight at 80 °C to obtain the MgAl-HT and ZnAl-HT.
Supported copper catalysts were synthesized through hydrothermal deposition precipitation method, with the theoretical loadings of 1 wt.%. At first, 2 g of the HT supports and 100 mL of deionized water were added into a round-bottom flask. By using the 0.16 M of Na2CO3, the pH value of the solution was adjusted to about 10. Then, 75 mg of Cu(NO3)2·3H2O was added into the flask and the solution was heated up to 80 °C. The mixture was stirred at 80 °C for 2 h, and the products were filtered and washed with deionized water. Before the test, the samples were dried at 80 °C overnight and calcined at 300 °C for 2 h. The obtained supported Cu catalysts were denoted as Cu/MgAlOx and Cu/ZnAlOx.

3.3. Catalytic Test

The catalytic conversion of ethanol into other products was carried out in a stainless steel fixed-bed reactor (inner diameter: 10 mm, length: 660 mm) equipped with a thermocouple and a mass flow controller. The reaction was performed at 280 °C with N2 as the carrier gas. Prior to the test, 1 g of the catalyst was loaded into the middle of the reactor, with quartz sand to transfer mass and heat. Then, the pressure of system was increased and maintained with a back-pressure regulator, with a gas hourly space velocity (GHSV) of 692 h−1. After purging with N2 for 0.5 h, the temperature was programmatically increased to 280 °C, with heating rate of 5 °C/min. Then, ethanol was introduced into the system by using a plunger pump (NP-KX-210) with liquid hourly space velocity (LHSV) of 4.8 h−1. The reaction products were analyzed offline with an Agilent 7890B chromatograph equipped with an HP-5 capillary column (30 m × 0.32 mm) and a flame ionization detector (FID). O-xylene was used as an internal standard for the quantification of the liquid products.
For the catalytic conversion of ethanol, the conversion was calculated as moles of ethanol reacted to the moles of ethanol fed to the reactor. The selectivity of each product was calculated as moles of carbon in the target product to the moles of carbon in the ethanol reacted. These calculations can be given by the following equations:
Conversion   ( % ) = Moles   of   ethanol   reacted Moles   of   ethanol   fed   to   the   reactor   ×   100
Selectivity   ( % ) = Moles   of   carbon   in   the   target   product Moles   of   carbon   in   ethanol   reacted   ×   100

3.4. Characterizations

The actual loadings of Cu for the catalysts were measured with an inductively coupled plasma atomic emission spectroscopy (ICP-AES) on an IRIS Intrepid II XSP instrument (Thermo Electron Corporation, Madison, WI, USA).
N2-physical adsorption-desorption tests were measured at 77 K using an AutoSorb-1 instrument. Prior to the measurements, the catalysts were treated at vacuum for 2 h at 120 °C. The pore size distribution and the specific surface area were calculated by the BJH and BET methods.
X-ray powder diffraction (XRD) patterns were conducted on a PW3040/60 X’Pert PRO (PANalytical, Almelo, Netherlands) diffractometer, equipped with a Cu Kα radiation source (λ = 0.15432 nm) operating at 40 kV and 40 mA.
Scanning transmission electron microscopy (STEM) measurements were conducted on a JEM-2100F microscope at 200 kV equipped with an energy dispersive X-ray (EDX) spectrometer. The samples were prepared by dispersing the catalyst powder in ethanol via ultra-sonication onto a micro molybdenum TEM grid.
The X-ray photoelectron spectra (XPS) were conducted on a Thermo ESCALAB 250 X-ray spectrometer (Thermo Fisher Scientific, Madison, WI, USA) equipped with a monochromated Al Kα anode. The binding energies were calibrated for surface charging by referencing them to the energy of the C 1s peak at 284.8 eV.
The temperature-programmed desorption of carbon dioxide (CO2-TPD) experiments were performed on an AutochemII 2920 instrument (Norcross, GA, USA) with a thermal conductivity detector and mass spectrometry. Before the test, 100 mg of the catalysts were added into a U-type quartz tube reactor. Then the samples were heated in a flow of helium at 300 °C for 30 min. After the temperature decreased to 100 °C, pulses of CO2 were introduced up to saturation of the sample. Then, the CO2-TPD signal were recorded from 100 °C to 900 °C at a rate of 10 °C /min with a cold trap. The signal of the desorbed CO2 was also recorded by the MS simultaneously.
Pyridine adsorbed Fourier-transformed infrared absorption spectra were conducted on a Bruker INVENIO spectrometer (Karlsruhe, Germany), equipped with a MCT detector in the range of 1000~4000 cm−1. Each spectrum was collected with a resolution of 4 cm−1 and 32 scans. Self-supported catalyst wafers (~0.15 g) were pressed and placed into a in situ IR cell with a CaF2 window. Before the test, the cell was vacuumed to 10−3 torr, and the catalyst was treated at 200 °C for 30 min with a heating ramp of 5 °C/min. After cooling to room temperature, the background spectrum was collected under vacuum condition at 25 °C, and then the sample was exposed to pyridine vapor until adsorption saturation. Finally, the adsorbed pyridine was desorbed at 100 °C until the spectra showed no change.
In situ diffuse reflectance infrared Fourier-transformed spectroscopy (DRIFTS) was also performed on a Bruker INVENIO spectrometer (Bruker, Karlsruhe, Germany). Prior to the test, the catalyst was packed into a in situ cell with a ZnSe window and treated at 300 °C for 0.5 h. After cooling to 25 °C, the background spectrum was recorded under atmospheric pressure in helium. Then, ethanol was introduced into the cell with the assistance of helium. Finally, the reaction temperature was programmatically increased to the fixed temperatures, with a heating rate of 10 °C/min. After achieving the steady state, the spectra were collected at 100, 150, 200, 250, and 300 °C, respectively.

4. Conclusions

In summary, MgAlOx and ZnAlOx mixed-oxides supported Cu catalysts were used for ethanol conversion via Guerbet coupling reaction. The two catalysts show obvious different performances under identical reaction conditions. For Cu/MgAlOx, conversion of ethanol and selectivity of butanol were 43.1% and 33.2%, respectively, whereas for Cu/ZnAlOx, the major product is ethyl acetate, with conversion and selectivity of 33.9% and 42.2%, respectively. The sole mixed-oxides were also conducted to make a comparison, from which low activity was observed in both catalysts, indicating a great influence of Cu species for ethanol transformation. The acid-base sites resulting from different mixed oxides might determine the distribution of products, from which the catalysts with appropriate basic sites were beneficial to form the condensation products such as crotonaldehyde and butanol, whereas the catalysts with more acid sites were believed to form dehydrogenation products, such as ethyl acetate. In combination the catalytic performances with the characterization results, the synergistic effect between Cu species with the acid-base sites were deduced to be the active sites. This work will provide good reference for designing supported metal catalyst for the Guerbet coupling process with good activity and high product selectivity.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal12101170/s1, Figure S1: XPS spectra of supported Cu catalysts after stability test; Figure S2: XRD patterns of supported Cu catalysts after stability test; Figure S3: TEM and HRTEM images of different supported Cu catalysts after stability test; Figure S4: EDX-mapping of Cu/MgAlOx catalyst after stability test.

Author Contributions

Y.T. and Y.D. conceived of the study; Z.L. and J.L. performed most of the experiments; L.G. contributed with some of the characterizations. All the authors contributed to the writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China under Grant No. 22102149.

Acknowledgments

The authors appreciate the support from the public testing platform of Zhejiang Normal University and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Thanks for the technical support with ICP, XRD, TEM, HRTEM, and XPS measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aitchison, H.; Wingad, R.L.; Wass, D.F. Homogeneous Ethanol to Butanol Catalysis-Guerbet Renewed. ACS Catal. 2016, 6, 7125–7132. [Google Scholar] [CrossRef] [Green Version]
  2. Angelici, C.; Weckhuysen, B.M.; Bruijnincx, P.C.A. Chemocatalytic Conversion of Ethanol into Butadiene and Other Bulk Chemicals. ChemSusChem 2013, 6, 1595–1614. [Google Scholar] [CrossRef] [PubMed]
  3. Phillips, S.D.; Jones, S.B.; Meyer, P.A.; Snowden-Swan, L.J. Techno-economic analysis of cellulosic ethanol conversion to fuel and chemicals. Biofuels Bioprod. Bioref. 2022, 16, 640–652. [Google Scholar] [CrossRef]
  4. Hamelinck, C.N.; Hooijdonk, G.; Faaij, A.P.C. Ethanol from lignocellulosic biomass: Techno-economic performance in short-, middle- and long-term. Biomass Bioenergy 2005, 28, 384–410. [Google Scholar] [CrossRef]
  5. Rass-Hansen, J.; Falsig, H.; Jørgensen, B.; Christensen, C.H. Perspective Bioethanol: Fuel or feedstock? Chem. Technol. Biotechnol. 2007, 82, 329–333. [Google Scholar] [CrossRef]
  6. Szulczyk, K.R. Which is a better transportation fuel-Butanol or ethanol? Int. J. Hydrog. Energy 2010, 3, 501–512. [Google Scholar]
  7. Dziugan, P.; Jastrzabek, K.G.; Binczarski, M.; Karski, S.; Witonska, I.A.; Kolesinska, B.; Kaminski, Z.J. Continuous catalytic coupling of raw bioethanol into butanol and higher homologues. Fuel 2015, 158, 81–90. [Google Scholar] [CrossRef]
  8. Carvalho, D.L.; Avillez, R.R.; Rodrigues, M.T.; Borges, L.E.P.; Appel, L.G. Mg and Al mixed oxides and the synthesis of n-butanol from ethanol. Appl. Catal. A Gen. 2012, 415–416, 96–100. [Google Scholar] [CrossRef]
  9. Chistyakov, A.V.; Nikolaev, S.A.; Zharova, P.A.; Tsodikov, M.V.; Manenti, F. Linear α-alcohols production from supercritical ethanol over Cu/Al2O3 catalyst. Energy 2019, 166, 569–576. [Google Scholar] [CrossRef]
  10. Pacheco, H.P.; Souza, E.F.; Landi, S.M.; David, M.V.; Prillaman, J.T.; Davis, R.J.; Toniolo, F.S. Ru Promoted MgO and Al-Modified MgO for Ethanol Upgrading. Top. Catal. 2019, 62, 894–907. [Google Scholar] [CrossRef]
  11. Osman, M.B.; Krafft, J.M.; Thomas, C.; Yoshioka, T.; Kubo, J.; Costentin, G. Importance of the Nature of the Active Acid/Base Pairs of Hydroxyapatite Involved in the Catalytic Transformation of Ethanol to n-Butanol Revealed by Operando DRIFTS. ChemCatChem 2019, 11, 1–15. [Google Scholar] [CrossRef]
  12. León, M.; Díaz, E.; Ordóñez, S. Ethanol catalytic condensation over Mg–Al mixed oxides derived from hydrotalcites. Catal. Today 2011, 164, 436–442. [Google Scholar] [CrossRef]
  13. Neumann, C.N.; Payne, M.T.; Rozeveld, S.J.; Wu, Z.; Zhang, G.; Comito, R.J.; Miller, J.T.; Dincă, M. Structural Evolution of MOF-Derived RuCo, A General Catalyst for the Guerbet Reaction. ACS Appl. Mater. Interfaces 2021, 13, 52113–52124. [Google Scholar] [CrossRef]
  14. Mück, J.; Kocík, J.; Hájek, M.; Tišler, Z.; Frolich, K.; Kašpárek, A. Transition metals promoting Mg-Al mixed oxides for conversion of ethanol to butanol and other valuable products: Reaction pathways. Appl. Catal. A Gen. 2021, 626, 118380. [Google Scholar] [CrossRef]
  15. López-Olmos, C.; Morales, M.V.; Guerrero-Ruiz, A.; Ramirez-Barria, C.; Asedegbega-Nieto, E.; Rodríguez-Ramos, I. Continuous gas phase condensation of bioethanol to 1-butanol over bifunctional Pd/Mg and Pd/Mg-carbon catalysts. ChemSusChem 2018, 11, 3502–3511. [Google Scholar] [CrossRef] [PubMed]
  16. Nikolaev, S.A.; Tsodikov, M.V.; Chistyakov, A.V.; Zharova, P.A.; Ezzgelenko, D.I. The activity of mono-and bimetallic gold catalysts in the conversion of sub-and supercritical ethanol to butanol. J. Catal. 2019, 369, 501–507. [Google Scholar] [CrossRef]
  17. Wu, X.; Fang, G.; Liang, Z.; Leng, W.; Xu, K.; Jiang, D.; Ni, J.; Li, X. Catalytic upgrading of ethanol to n-butanol over M-CeO2/AC (M=Cu, Fe, Co, Ni and Pd) catalysts. Catal. Commun. 2017, 100, 15–18. [Google Scholar] [CrossRef]
  18. Fu, S.; Shao, Z.; Wang, Y.; Liu, Q. Manganese-Catalyzed Upgrading of Ethanol into 1-Butanol. J. Am. Chem. Soc. 2017, 139, 11941–11948. [Google Scholar] [CrossRef]
  19. Perrone, O.M.; Lobefaro, F.; Aresta, M.; Nocito, F.; Boscolo, M.; Dibenedetto, A. Butanol synthesis from ethanol over CuMgAl mixed oxides modified with palladium (II) and indium (III). Fuel Process. Technol. 2018, 177, 353–357. [Google Scholar] [CrossRef] [Green Version]
  20. Hanspal, S.; Young, Z.D.; Prillaman, J.T.; Davis, R.J. Influence of surface acid and base sites on the Guerbet coupling of ethanol to butanol over metal phosphate catalysts. J. Catal. 2017, 352, 182–190. [Google Scholar] [CrossRef]
  21. Tsuchida, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Direct Synthesis of n-Butanol from Ethanol over Nonstoichiometric Hydroxyapatite. Ind. Eng. Chem. Res. 2006, 45, 8634–8642. [Google Scholar] [CrossRef]
  22. Ogo, S.; Onda, A.; Yanagisawa, K. Selective synthesis of 1-butanol from ethanol over strontium phosphate hydroxyapatite catalysts. Appl. Catal. A Gen. 2011, 402, 188–195. [Google Scholar] [CrossRef]
  23. IMarcu, C.; Tichit, D.; Fajula, F.; Tanchoux, N. Catalytic valorization of bioethanol over Cu-Mg-Al mixed oxide catalysts. Catal. Today 2019, 147, 231–238. [Google Scholar]
  24. Jiang, D.; Wu, X.; Mao, J.; Ni, J.; Li, X. Continuous catalytic upgrading of ethanol to n-butanol over Cu–CeO2/AC catalysts. Chem. Commun. 2016, 52, 13749. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, Z.; Pang, J.; Song, L.; Li, X.; Yuan, Q.; Li, X.; Liu, S.; Zheng, M. Conversion of Ethanol to n-Butanol over NiCeO2 Based Catalysts: Effects of Metal Dispersion and NiCe Interactions. Ind. Eng. Chem. Res. 2020, 59, 22057–22067. [Google Scholar] [CrossRef]
  26. Yang, C.; Meng, Z.Y. Bimolecular Condensation of Ethanol to 1-Butanol Catalyzed by Alkali Cation Zeolites. J. Catal. 1993, 142, 37–44. [Google Scholar] [CrossRef]
  27. Wang, Z.; Yin, M.; Pang, J.; Li, X.; Xing, Y.; Su, Y.; Liu, S.; Liu, X.; Wu, P.; Zheng, M.; et al. Active and stable Cu doped NiMgAlO catalysts for upgrading ethanol to n-butanol. J. Energy Chem. 2022, 72, 306–317. [Google Scholar] [CrossRef]
  28. Cuello-Penaloza, P.A.; Dastidar, R.G.; Wang, S.C.; Du, Y.; Lanci, M.P.; Wooler, B.; Kliewer, C.E.; Hermans, I.; Dumesic, J.A.; Huber, G.W. Ethanol to distillate-range molecules using Cu/MgxAlOy catalysts with low Cu loadings. Appl. Catal. B Environ. 2022, 304, 120984. [Google Scholar] [CrossRef]
  29. Zhou, J.; He, Y.; Xue, B.; Cheng, Y.; Zhou, D.; Wang, D.; He, Y.; Guan, W.; Fang, K.; Zhang, L.; et al. Benefits of active site proximity in Cu@UiO-66 catalysts for efficient upgrading of ethanol to nbutanol. Sustain. Energy Fuels 2021, 5, 4628. [Google Scholar] [CrossRef]
  30. Jiang, D.; Fang, G.; Tong, Y.; Wu, X.; Wang, Y.; Hong, D.; Leng, W.; Liang, Z.; Tu, P.; Liu, L.; et al. Multifunctional Pd@UiO-66 Catalysts for Continuous Catalytic Upgrading of Ethanol to n-Butanol. ACS Catal. 2018, 8, 11973–11978. [Google Scholar] [CrossRef]
  31. Li, J.; Lin, L.; Tan, Y.; Wang, S.; Yang, W.; Chen, X.; Luo, W.; Ding, Y. High performing and stable Cu/NiAlOx catalysts for the continuous catalytic conversion of ethanol into butanol. ChemCatChem 2022, 14, e202200539. [Google Scholar]
  32. Gao, D.; Feng, Y.; Yin, H.; Wang, A.; Jiang, T. Coupling reaction between ethanol dehydrogenation and maleic anhydride hydrogenation catalyzed by Cu/Al2O3, Cu/ZrO2, and Cu/ZnO catalysts. Chem. Eng. J. 2013, 233, 349–359. [Google Scholar] [CrossRef]
  33. Cheng, F.; Guo, H.; Cui, J.; Hou, B.; Li, D. Guerbet reaction of methanol and ethanol catalyzed by CuMgAlOx mixed oxides: Effect of M2+/Al3+ ratio. J. Fuel Chem. Technol. 2018, 46, 1472–1481. [Google Scholar] [CrossRef]
  34. Petrolini, D.D.; Eagan, N.; Ball, M.R.; Burt, S.P.; Hermans, I.; Huber, G.W.; Dumesic, J.A.; Martins, L. Ethanol condensation at elevated pressure over copper on AlMgO and AlCaO porous mixed-oxide supports. Catal. Sci. Technol. 2019, 9, 2032. [Google Scholar] [CrossRef]
  35. Zhu, Q.; Yin, L.; Ji, K.; Li, C.; Wang, B.; Tan, T. Effect of Catalyst Structure and Acid−Base Property on the Multiproduct Upgrade of Ethanol and Acetaldehyde to C4 (Butadiene and Butanol) over the Y−SiO2 Catalysts. ACS Sustain. Chem. Eng. 2020, 8, 1555–1565. [Google Scholar] [CrossRef]
  36. Sing, K.S.W.; Williams, R.T. Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials. Adsorpt. Sci. Technol. 2004, 22, 773–782. [Google Scholar] [CrossRef]
  37. Martunus; Othman, M.R.; Fernando, W.J.N. Elevated temperature carbon dioxide capture via reinforced metal hydrotalcite. Micropor. Mesopor. Mater. 2011, 138, 110–117. [Google Scholar] [CrossRef]
  38. Sutthiumporn, K.; Kawi, S. Promotional effect of alkaline earth over Ni-La2O3 catalyst for CO2 reforming of CH4: Role of surface oxygen species on H2 production and carbon suppression. Int. J. Hydrog. Energ. 2011, 36, 14435–14446. [Google Scholar] [CrossRef]
  39. Pang, J.; Zheng, M.; Wang, C.; Yang, X.; Liu, H.; Liu, X.; Sun, J.; Wang, Y.; Zhang, T. Hierarchical Echinus-like Cu-MFI Catalysts for Ethanol Dehydrogenation. ACS Catal. 2020, 10, 13624–13629. [Google Scholar] [CrossRef]
  40. Wang, L.; Zhu, W.; Zheng, D.; Yu, X.; Cui, J.; Jia, M.; Zhang, W.; Wang, Z. Direct transformation of ethanol to ethyl acetate on Cu/ZrO2 catalyst. Reac. Kinet. Mech. Cat. 2010, 101, 365–375. [Google Scholar] [CrossRef]
  41. Zhang, J.; Shi, K.; Zhu, Y.; An, Z.; Wang, W.; Ma, X.; Shu, X.; Song, H.; Xiang, X.; He, J. Interfacial Sites in Ag Supported Layered Double Oxide for Dehydrogenation Coupling of Ethanol to n-Butanol. ChemistryOpen 2021, 10, 1095–1103. [Google Scholar] [CrossRef] [PubMed]
  42. Li, H.; Tan, Y.; Chen, X.; Yang, W.; Huang, C.; Li, J.; Ding, Y. Efficient Synthesis of Methyl Methacrylate by One Step Oxidative Esterification over Zn-Al-Mixed Oxides Supported Gold Nanocatalysts. Catalysts 2021, 11, 162. [Google Scholar] [CrossRef]
  43. Gao, J.; Fan, G.; Yang, L.; Cao, X.; Zhang, P.; Li, F. Oxidative Esterification of Methacrolein to Methyl Methacrylate over Gold Nanoparticles on Hydroxyapatite. ChemCatChem 2017, 9, 1230–1241. [Google Scholar] [CrossRef]
  44. Liang, Z.; Jiang, D.; Fang, G.; Leng, W.; Tu, P.; Tong, Y.; Liu, L.; Ni, J.; Li, X. Catalytic Enhancement of Aldol Condensation by Oxygen Vacancy on CeO2 Catalysts. ChemistrySelect 2019, 4, 4364–4370. [Google Scholar] [CrossRef]
  45. Zhang, J.; Shi, K.; An, Z.; Zhu, Y.; Shu, X.; Song, H.; Xiang, X.; He, J. Acid−Base Promoted Dehydrogenation Coupling of Ethanol on Supported Ag Particles. Ind. Eng. Chem. Res. 2020, 59, 3342–3350. [Google Scholar] [CrossRef]
  46. Pinzón, M.; Cortés-Reyes, M.; Herrera, C.; Larrubia, M.Á.; Alemany, L.J. Ca-based bifunctional acid-basic model-catalysts for n-butanol production from ethanol condensation. Biofuels Bioprod. Bior. 2021, 15, 218–230. [Google Scholar] [CrossRef]
  47. Yuan, B.; Zhang, J.; An, Z.; Zhu, Y.; Shu, X.; Song, H.; Xiang, X.; Wang, W.; Jing, Y.; Zheng, L.; et al. Atomic Ru catalysis for ethanol coupling to C4+ alcohols. Appl. Catal. B 2022, 309, 121271. [Google Scholar] [CrossRef]
  48. Tong, Y.; Zhou, J.; He, Y.; Tu, P.; Xue, B.; Cheng, Y.; Cen, J.; Zheng, Y.; Ni, J.; Li, X. Structure-activity relationship of Cu species in the ethanol upgrading to n-butanol. ChemistrySelect 2020, 5, 7714–7719. [Google Scholar] [CrossRef]
  49. Riittonen, T.; Toukoniitty, E.; Madnani, D.K.; Leino, A.R.; Kordas, K.; Szabo, M.; Sapi, A.; Arve, K.; Wärnå, J.; Mikkola, J.P. One-pot liquid-phase catalytic conversion of ethanol to 1-butanol over aluminium oxide—The effect of the active metal on the selectivity. Catalysts 2012, 2, 68–84. [Google Scholar] [CrossRef]
Figure 1. The products distributions of reaction with time courses for the (a) Cu/MgAlOx and (b) Cu/ZnAlOx catalysts. Reaction conditions: amounts of catalyst: 1 g; pressure of N2: 2 MPa; temperature: 553K; GHSV: 692 h−1, LHSV: 4.8 h−1.
Figure 1. The products distributions of reaction with time courses for the (a) Cu/MgAlOx and (b) Cu/ZnAlOx catalysts. Reaction conditions: amounts of catalyst: 1 g; pressure of N2: 2 MPa; temperature: 553K; GHSV: 692 h−1, LHSV: 4.8 h−1.
Catalysts 12 01170 g001
Figure 2. (a) N2 physical adsorption-desorption isotherms and (b) BJH-pore size distributions of mixed oxides and supported Cu catalysts.
Figure 2. (a) N2 physical adsorption-desorption isotherms and (b) BJH-pore size distributions of mixed oxides and supported Cu catalysts.
Catalysts 12 01170 g002
Figure 3. XRD patterns of different supported Cu catalysts.
Figure 3. XRD patterns of different supported Cu catalysts.
Catalysts 12 01170 g003
Figure 4. TEM and HRTEM images of different supported Cu catalysts: (a,c) Cu/MgAlOx; (b,d) Cu/ZnAlOx.
Figure 4. TEM and HRTEM images of different supported Cu catalysts: (a,c) Cu/MgAlOx; (b,d) Cu/ZnAlOx.
Catalysts 12 01170 g004
Figure 5. XPS spectra of different supported Cu catalysts: (a) C 1s; (b) O 1s; (c) Cu 2p; and (d) Al 2p.
Figure 5. XPS spectra of different supported Cu catalysts: (a) C 1s; (b) O 1s; (c) Cu 2p; and (d) Al 2p.
Catalysts 12 01170 g005
Figure 6. (a) CO2 TPD profile and (b) pyridine adsorbed FT-IR spectra of the supported Cu catalysts.
Figure 6. (a) CO2 TPD profile and (b) pyridine adsorbed FT-IR spectra of the supported Cu catalysts.
Catalysts 12 01170 g006
Figure 7. In situ drift spectra of ethanol adsorption and transformation on (a) Cu/MgAlOx and (b) Cu/ZnAlOx catalysts.
Figure 7. In situ drift spectra of ethanol adsorption and transformation on (a) Cu/MgAlOx and (b) Cu/ZnAlOx catalysts.
Catalysts 12 01170 g007
Scheme 1. The proposed mechanisms for synthesis of (a) butanol and (b) ethyl acetate for the Cu/MgAlOx and Cu/ZnAlOx catalysts.
Scheme 1. The proposed mechanisms for synthesis of (a) butanol and (b) ethyl acetate for the Cu/MgAlOx and Cu/ZnAlOx catalysts.
Catalysts 12 01170 sch001
Table 1. Catalytic activities of mixed oxides and supported Cu catalysts for ethanol conversion.
Table 1. Catalytic activities of mixed oxides and supported Cu catalysts for ethanol conversion.
EntryCatalystsConversion (%) Selectivity (%)
ButanolEthyl AcetateEtherButaldehydeEthyl ButyrateHexanol
1MgAlOx4.400.71.51.500
2ZnAlOx8.50.238.602.11.50
3Cu/MgAlOx43.133.211.808.65.08.0
4Cu/ZnAlOx33.97.742.21.89.48.30
Reaction conditions: Catalyst: 1 g, T = 553 K, PN2 = 2 MPa, GHSV = 692 h−1, LHSV = 4.8 h−1.
Table 2. Textural properties of mixed oxides and supported Cu catalysts.
Table 2. Textural properties of mixed oxides and supported Cu catalysts.
EntryCatalystLoadings of Cu
(%)
Surface Area
(m2/g)
Pore Volume
(cm3/g)
Half Pore Width
(nm)
1MgAlOx-119.70.4412.69
2ZnAlOx-155.90.303.43
3Cu/MgAlOx1.11101.50.364.87
4Cu/ZnAlOx1.35115.70.251.71
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Z.; Li, J.; Tan, Y.; Guo, L.; Ding, Y. Copper Supported on MgAlOx and ZnAlOx Porous Mixed-Oxides for Conversion of Bioethanol via Guerbet Coupling Reaction. Catalysts 2022, 12, 1170. https://doi.org/10.3390/catal12101170

AMA Style

Liu Z, Li J, Tan Y, Guo L, Ding Y. Copper Supported on MgAlOx and ZnAlOx Porous Mixed-Oxides for Conversion of Bioethanol via Guerbet Coupling Reaction. Catalysts. 2022; 12(10):1170. https://doi.org/10.3390/catal12101170

Chicago/Turabian Style

Liu, Zongyang, Jie Li, Yuan Tan, Luyao Guo, and Yunjie Ding. 2022. "Copper Supported on MgAlOx and ZnAlOx Porous Mixed-Oxides for Conversion of Bioethanol via Guerbet Coupling Reaction" Catalysts 12, no. 10: 1170. https://doi.org/10.3390/catal12101170

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop